首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 578 毫秒
1.
The kinetics for the reactions of C6H5 with phenylacetylene and styrene have been measured by CRDS in the temperature range 297-409 K under an Ar pressure of 3.6 Torr. The total rate constants can be given by the following Arrhenius expressions (in units of cm3 mol−1 s−1): k1(C6H5 + C6H5C2H) = 1013.0±0.1exp [−(2430 ± 150)/RT] and k2(C6H5 + C6H5C2H3) = 1013.3±0.1 exp [−(2570 ± 180)/T]. Additional DFT and MP2 calculations have been carried out to assist our interpretation of the measured kinetic data. The addition of C6H5 to the terminal CHx (x = 1 or 2) sites is predicted to be the dominant channel for both reactions. The calculated bimolecular rate constants are in reasonable agreement with experimental values for the temperature range studied.  相似文献   

2.
Kinetics and mechanisms for reactions of OH with methanol and ethanol have been investigated at the CCSD(T)/6-311 + G(3df2p)//MP2/6-311 + G(3df2p) level of theory. The total and individual rate constants, and product branching ratios for the reactions have been computed in the temperature range 200-3000 K with variational transition state theory by including the effects of multiple reflections above the wells of their pre-reaction complexes, quantum-mechanical tunneling and hindered internal rotations. The predicted results can be represented by the expressions k1 = 4.65 × 10−20 × T2.68 exp(414/T) and k2 = 9.11 × 10−20 × T2.58 exp(748/T) cm3 molecule−1 s−1 for the CH3OH and C2H5OH reactions, respectively. These results are in reasonable agreements with available experimental data except that of OH + C2H5OH in the high temperature range. The former reaction produces 96-89% of the H2O + CH2OH products, whereas the latter process produces 98-70% of H2O + CH3CHOH and 2-21% of the H2O + CH2CH2OH products in the temperature range computed (200-3000 K).  相似文献   

3.
The effects of temperature and pressure on the formation and decomposition of C6H5C2H2O2 in the C6H5C2H2 + O2 reaction have been investigated at temperatures from 298 to 378 K by directly monitoring the C6H5C2H2O2 radical in the visible region by cavity ringdown spectrometry (CRDS). The rate constant for the C6H5C2H2 + O2 association and that for fragmentation of C6H5C2H2O2 were found to be k1 (C6H5C2H2 + O2 → C6H5C2H2O2) = (3.20 ± 1.19) × 1011 exp(+760/T) cm3 mol−1 s−1 and k2 (C6H5C2H2 O2 → C6H5CHO + HCO) = (1.68 ± 0.13) × 104 s−1, respectively. Additional kinetic measurements by pulsed laser photolysis/mass spectrometry show that C6H5CHO was produced in the C6H5C2H2 + O2 reaction as predicted and the formation of C6H5CHO from the decomposition of C6H5C2H2O2 is temperature-independent, consistent with the CRDS experimental data.  相似文献   

4.
Single-pulse shock-tube experiments were used to study the thermal decomposition of selected oxygenated hydrocarbons: Ethyl propanoate (C2H5OC(O)C2H5; EP), propyl propanoate (C3H7OC(O)C2H5; PP), isopropyl acetate ((CH3)2HCOC(O)CH3; IPA), and methyl isopropyl carbonate ((CH3)2HCOC(O)OCH3; MIC) The consumption of reactants and the formation of stable products such as C2H4 and C3H6 were measured with gas chromatography/mass spectrometry (GC/MS). Depending on the considered reactant, the temperatures range from 716–1102 K at pressures between 1.5 and 2.0 bar. Rate-coefficient data were obtained from first-order analysis. All reactants primarily decompose by six-center eliminations: EP → C2H4 + C2H5COOH (propionic acid); PP → C3H6 + C2H5COOH; IPA → C3H6 + CH3COOH (acetic acid); MIC → C3H6 + CH3OC(O)OH (methoxy formic acid). Experimental rate-coefficient data can be well represented by the following Arrhenius expressions: k(EP → products) = 1013.49±0.16 exp(−214.95±3.25 kJ/mol/RT) s−1; k(PP → products) = 1012.21±0.16 exp(–191.21±2.79 kJ/mol/RT) s−1; k(IPA → products) = 1013.10±0.31 exp(–186.38±5.10 kJ/mol/RT) s−1; k(MIC → products) = 1012.43±0.29 exp(–165.25±4.46 kJ/mol/RT) s−1. The determination of rate coefficients was based on the amount of C2H4 or C3H6 formed. The potential energy surface (PES) of the thermal decomposition of these four reactants was determined with the G4 composite method. A master-equation analysis was conducted based on energies and molecular properties from the G4 computations. The results indicate that the length of a linear alkyl substituent does not significantly influence the rate of six-center eliminations, whereas the change from a linear to a branched alkyl substituent results in a significant reactivity increase. The comparison between rate-coefficient data also shows that alkyl carbonates have higher reactivity towards decomposition by six-center elimination than esters. The results are discussed in in the context of reactivity patterns of carbonyl compounds.  相似文献   

5.
This work reports measurements of the absolute rate coefficients and Rice-Ramsperger-Kassel-Markus (RRKM) master equation (ME) simulations of the C2H3 + C3H6 reaction. Direct kinetic studies were performed over a temperature range of 300-700 K and pressures of 15, 25, and 100 Torr. Vinyl radicals were generated by laser photolysis of vinyl iodide at 266 nm, and time-resolved absorption spectroscopy was used to probe vinyl radicals through absorption at 423.2 nm. A weighted modified Arrhenius fit to the experimental rate constant is k1 = (1.3 ± 0.2) × 10−12 cm3 molecule−1 s−1(T/1000)1.6 exp[−(1510 ± 80/T)]. Fifteen stationary points and 48 transition states on the C5H9 potential energy surface (PES) were calculated using the G3 method in Gaussian 03. RRKM/ME simulations were performed using VariFlex on a simplified PES to predict pressure dependent rate coefficients and branching fractions for the major channels. For temperatures between 350 and 700 K, the calculated rate coefficient agrees with the experimental rate coefficient within 20%. At low temperatures, the primary products are the initial adducts 4-penten-2-yl and 2-methyl-3-buten-1-yl. At higher temperatures, the dominant products are 1,3-butadiene + methyl, allyl + ethene, and 1,3-pentadiene + H. Although C2H3 + C3H6 → allyl + ethene is thermodynamically favored, the simulations predict that it does not become the dominant product until 1700 K.  相似文献   

6.
The primary product formation of the C3H5 + O reaction in the gas phase has been studied at room temperature. Allyl radicals (C3H5) and O atoms were generated by laser flash photolysis at λ = 193 nm of the precursors C3H5Cl, C3H5Br, C6H10 (1,5-hexadiene), and SO2, respectively. The educts and the products were detected by using quantitative FTIR spectroscopy. The combined product analysis of the experiments with the different precursors leads to the following relative branching fractions: C3H5 + O → C3H4O + H (47%), C2H4 + H + CO (41%), H2CO + C2H2 + H (7%), CH3CCH + OH and CH2CCH2 + OH (<5%). The rate of reaction has been studied relative to CH3OCH2 + O and C2H5 + O in the temperature range from 300 to 623 K. Here, the radicals were produced via the fast reactions of propene, dimethyl ether, and ethane, respectively, with atomic fluorine. Laser-induced multiphoton ionization combined with TOF mass spectrometry and molecular beam sampling from a flow reactor was used for the specific and sensitive detection of the C3H5, C2H5, and CH3COCH2 radicals. The rate coefficient of the reaction C3H5 + O was derived with reference to the reaction C2H5 + O leading to k(C3H5 + O) = (1.11 ± 0.2) × 1014 cm3/(mol s) in the temperature range 300-623 K. The C3H5 + O rate and channel branching, when incorporated in a suitable detailed reaction mechanism, have a large influence on benzene and allyl concentration profiles in fuel-rich propene flames, on the propene flame speed, and on propene ignition delay times.  相似文献   

7.
The temporal variation of chemiluminescence emission from OH?(A2 Σ +) and CH?(A2 Δ) in reacting Ar-diluted H2/O2/CH4, C2H2/O2 and C2H2/N2O mixtures was studied in a shock tube for a wide temperature range at atmospheric pressures and various equivalence ratios. Time-resolved emission measurements were used to evaluate the relative importance of different reaction pathways. The main formation channel for OH? in hydrocarbon combustion was studied with CH4 as benchmark fuel. Three reaction pathways leading to CH? were studied with C2H2 as fuel. Based on well-validated ground-state chemistry models from literature, sub-mechanisms for OH? and CH? were developed. For the main OH?-forming reaction CH+O2=OH?+CO, a rate coefficient of k 2=(8.0±2.6)×1010 cm3?mol?1?s?1 was determined. For CH? formation, best agreement was achieved when incorporating reactions C2+OH=CH?+CO (k 5=2.0×1014 cm3?mol?1?s?1) and C2H+O=CH?+CO (k 6=3.6×1012exp(?10.9 kJ?mol?1/RT) cm3?mol?1?s?1) and neglecting the C2H+O2=CH?+CO2 reaction.  相似文献   

8.
The high-temperature photochemistry (HTP) technique, previously used for reactions of neutral species, has been adapted to the study of atomic metal ion-molecule reactions. Ca+ ions were generated by 193 nm multi-photon photolysis of calcium acetyl acetonate and its pyrolysis fragments. The relative ion concentrations were monitored by laser-induced fluorescence at 393.4 nm. Ar was used as the bath gas. The data for the Ca+ + O2 + M → CaO2+ + M association reaction (1) are fitted by k1(907-1425 K) = 3.5 × 10−32 exp(+3161 K/T) cm6 molecule−2 s−1. Combining with an approximate k1(296 K) value in the literature leads to k1(296-1425 K) = 5.8 × 10−22 (T/K)−2.9 exp(−601 K/T) cm6 molecule−2 s−1. Over much of the observed temperature range reaction (1) has much smaller rate coefficients than the corresponding neutral Ca association reaction. Reaction (1) is shown to behave very similarly to the O2 association reaction with neutral K atoms, with which Ca+ is iso-electronic. This suggests that the initial step is ion-pair complex formation of the superoxide Ca2+(O2), which is also consistent with results from density functional calculations. The k1 values are rationalized via Troe’s unimolecular formalism, which leads to good accord with the experiments.  相似文献   

9.
The water-soluble Pr (Ⅲ) and Nd (Ⅲ) complexes with an ofloxacin derivative have been prepared and characterized. The single-crystal X-ray diffraction showed that the Pr (III) and Nd (III) complexes have the similar molecular structure. Under physiological pH condition, the effects of [PrL(NO3)2(CH3OH)](NO3) and [NdL(NO3)2(CH3OH)](NO3) on bovine serum albumin (BSA) were examined using fluorescence spectroscopy in combination with UV-vis absorbance and circular dichroism (CD) spectra. The result reveals that the quenching mechanism of fluorescence of BSA by two complexes is a static quenching process and the number of binding sites is about 1 for both. The thermodynamic parameters (ΔH=−14.52 kJ mol−1, ΔS=56.54 J mol−1 K−1 for [PrL(NO3)2(CH3OH)](NO3) and ΔH=−24.63 kJ mol−1, ΔS=22.07 J mol−1 K−1 for [NdL(NO3)2(CH3OH)](NO3)) indicate that hydrophobic and electrostatic interactions are the main binding force in the complexes-BSA system. The binding average distance between complexes and BSA was obtained on the basis of Förster's theory. In addition, it was proved by the CD spectra that the BSA secondary structure was changed in the presence of complexes in an aqueous solution.  相似文献   

10.
a-C:H films were prepared by middle frequency plasma chemical vapor deposition (MF-PCVD) on silicon substrates from two hydrocarbon source gases, CH4 and a mixture of C2H2 + H2, at varying bias voltage amplitudes. Raman spectroscopy shows that the structure of the a-C:H films deposited from these two precursors is different. For the films deposited from CH4, the G peak position around 1520 cm−1 and the small intensity ratio of D peak to G peak (I(D)/I(G)) indicate that the C-C sp3 fraction in this film is about 20 at.%. These films are diamond-like a-C:H films. For the films deposited from C2H2 + H2, the Raman results indicate that their structure is close to graphite-like amorphous carbon. The hardness and elastic modulus of the films deposited from CH4 increase with increasing bias voltage, while a decrease of hardness and elastic modulus of the films deposited from a mixture of C2H2 + H2 with increasing bias voltage is observed.  相似文献   

11.
Single crystals of organic nonlinear optical (NLO) materials l-Histidine nitrate (C6H10N3O2)+ · (NO3) and l-Cysteine tartrate monohydrate (C3H8NO2S)+ · (C4H5O6) · H2O were grown by submerged seed solution method. Characterization of the crystals was made using single crystal X-ray diffraction. Fourier transform infrared (FTIR) spectroscopic studies, optical behaviour such as UV-visible-NIR absorption spectra and second harmonic generation (SHG) conversion efficiency were investigated to explore the NLO characteristics of the above materials. Microhardness measurements and dielectric studies of the compounds were also carried out.  相似文献   

12.
Laminar flame speeds were accurately measured for CO/H2/air and CO/H2/O2/helium mixtures at different equivalence ratios and mixing ratios by the constant-pressure spherical flame technique for pressures up to 40 atmospheres. A kinetic mechanism based on recently published reaction rate constants is presented to model these measured laminar flame speeds as well as a limited set of other experimental data. The reaction rate constant of CO + HO2 → CO2 + OH was determined to be k = 1.15 × 105T2.278 exp(−17.55 kcal/RT) cm3 mol−1 s−1 at 300-2500 K by ab initio calculations. The kinetic model accurately predicts our measured flame speeds and the non-premixed counterflow ignition temperatures determined in our previous study, as well as homogeneous system data from literature, such as concentration profiles from flow reactor and ignition delay time from shock tube experiments.  相似文献   

13.
Chemical preparation, calorimetric studies, crystal structure and spectroscopic investigations are given for a new noncentrosymmetric organic cation monophosphate [2,5-(CH3)2C6H3NH3]H2PO4. This compound is orthorhombic P212121 with the following unit-cell parameters: a=5.872(4), b=20.984(3), c=8.465(1) Å, Z=4, V=1043.0(5) Å3 and Dx=1.396 g cm−3. Crystal structure has been solved and refined to R=0.048 using 2526 independent reflections. Structure can be described as an inorganic layer parallel to (a,b) planes between which organic groups [2,5-(CH3)2C6H3NH3]+ are located. Multiple hydrogen bonds connecting the different entities of compound thrust upon three-dimensional network a noncentrosymmetric configuration.  相似文献   

14.
The C7H7 potential energy surface was studied from first principles to determine the benzyl radical decomposition mechanism. The investigated high temperature reaction pathway involves 15 accessible energy wells connected by 25 transition states. The analysis of the potential energy surface, performed determining kinetic constants of each elementary reaction using conventional transition state theory, evidenced that the reaction mechanism has as rate determining step the isomerization of the 1,3-cyclopentadiene, 5-vinyl radical to the 2-cyclopentene,5-ethenylidene radical and that the fastest reaction channel is dissociation to fulvenallene and hydrogen. This is in agreement with the literature evidences reporting that benzyl decomposes to hydrogen and a C7H6 species. The benzyl high-pressure decomposition rate constant estimated assuming equilibrium between the rate determining step transition state and benzyl is k1(T) = 1.44 × 1013T0.453exp(−38400/T) s−1, in good agreement with the literature data. As fulvenallene reactivity is mostly unknown, we investigated its reaction with hydrogen, which has been proposed in the literature as a possible decomposition route. The reaction proceeds fast both backward to form again benzyl and, if hydrogen adds to allene, forward toward the decomposition into the cyclopentadienyl radical and acetylene with high-pressure kinetic constants k2(T) = 8.82 × 108T1.20exp(1016/T) and k3(T) = 1.06 × 108T1.35exp(1716/T) cm3/mol/s, respectively. The computed rate constants were then inserted in a detailed kinetic mechanism and used to simulate shock tube literature experiments.  相似文献   

15.
The dissociation of acetone: CH3COCH3 → CH3CO + CH3, quickly followed by CH3CO → CH3 + CO, has been examined with Laser-Schlieren measurements in incident shock waves over 32-717 Torr and 1429-1936 K using 5% acetone dilute in krypton. A few very low pressure experiments (∼10 Torr) were used in a marginal effort to resolve the extremely fast vibrational relaxation of this molecule. This effort was partly motivated as a test for molecular, “roaming methyl” reactions, and also as a source of methyl radicals to test the application of a recent high-temperature mechanism for ethane decomposition [J.H. Kiefer, S. Santhanam, N.K. Srinivasan, R.S. Tranter, S.J. Klippenstein, M.A. Oehlschlaeger, Proc. Combust. Inst. 30 (2005) 1129-1135] on the reverse methyl combination. The gradient profiles show strong initial positive gradients and following negative values fully consistent with methyl radical formation and its following recombination. Thus C-C fission is certainly a large part of the process and molecular channels cannot be responsible for more than 30% of the dissociation. Rates obtained for the C-C fission show strong falloff well fit by variable reaction coordinate transition state theory when combined with a master equation. The calculated barrier is 82.8 kcal/mol, the fitted 〈ΔEdown = 400 (T/298) cm−1, similar to what was found in a recent study of C-C fission in acetaldehyde, and the extrapolated k = 1025.86 T−2.72 exp(−87.7 (kcal/mol)/RT), which agrees with the literature rate for CH3 + CH3CO. Large negative (exothermic) gradients appearing late from methyl combination are accurately fit in both time of onset and magnitude by the earlier ethane dissociation mechanism. The measured dissociation rates are in close accord with one earlier shock-tube study [K. Sato, Y. Hidaka, Combust. Flame 122 (2000) 291-311], but show much less falloff than the high pressure experiments of Ernst et al. [J. Ernst, K. Spindler, H.Gg. Wagner, Ber. Bunsenges. Phys. Chem. 80 (1976) 645-650].  相似文献   

16.
In the process of investigating the interaction of fullerene projectiles with adsorbed organic layers, we measured the kinetic energy distributions (KEDs) of fragment and parent ions sputtered from an overlayer of polystyrene (PS) oligomers cast on silver under 15 keV C60+ bombardment. These measurements have been conducted using our TRIFT™ spectrometer, recently equipped with the C60+ source developed by Ionoptika, Ltd. For atomic ions, the intensity corresponding to the high energy tail decreases in the following order: C+(E−0.4) > H+(E−1.5) > Ag+(E−3.5). In particular, the distribution of Ag+ is not broader than those of Ag2+ and Ag3+ clusters, in sharp contrast with 15 keV Ga+ bombardment. On the other hand, molecular ions (fragments and parent-like species) exhibit a significantly wider distribution using C60+ instead of Ga+ as primary ions. For instance, the KED of Ag-cationized PS oligomers resembles that of Ag+ and Agn+ clusters. A specific feature of fullerene projectiles is that they induce the direct desorption of positively charged oligomers, without the need of a cationizing metal atom. The energy spectrum of these PS+ ions is significantly narrower then that of Ag-cationized oligomers. For characteristic fragments of PS, such as C7H7+ and C15H13+ and polycyclic fragments, such as C9H7+ and C14H10+, the high energy decay is steep (E−4 − E−8). In addition, reorganized ions generally show more pronounced high energy tails than characteristic ions, similar to the case of monoatomic ion bombardment. This observation is consistent with the higher excitation energy needed for their formation. Finally, the fraction of hydrocarbon ions formed in the gas phase via unimolecular dissociation of larger species is slightly larger with gallium than with fullerene projectiles.  相似文献   

17.
Hai Hua Tang 《Surface science》2007,601(16):3293-3302
The interaction of ethyl vinyl ketone (EVK) with Si(1 1 1)-7 × 7 has been investigated using high-resolution electron energy loss spectroscopy (HREELS), X-ray photoelectron spectroscopy (XPS) and density functional theory (DFT) calculations. The disappearance of both stretching vibrations of CH2 (3099 cm−1) and CO (1684 cm−1) coupled with the appearance of new CC stretching mode (1660 cm−1) in the HREELS spectra of chemisorbed EVK clearly demonstrates the direct involvement of conjugated CC and CO bonds to form a SiC1H2C2HC3(C4H2C5H3)OSi surface species via [4 + 2]-like cycloaddition in a highly selective manner. In addition, XPS studies show that the C1s binding energies of C1/C2 and C3 upon chemisorption display chemical downshifts of 0.8 eV and 2.2 eV, respectively, further confirming the proposed [4 + 2]-like cycloaddition reaction for the EVK/Si(1 1 1)-7 × 7 system. DFT theoretical calculations suggest that the proposed [4 + 2]-like cycloadduct is thermodynamically most favorable.  相似文献   

18.
The adsorption of ethylene on Cu12Pt2 clusters has been studied within the density functional theory (DFT) approach to understand the high ethylene selectivity of Cu-rich Pt-Cu catalyst particles in the reaction of hydrogen-assisted 1,2-dichloroethane dechlorination. The structural parameters for Cu12Pt2 clusters with D4h, D2d, and C3v symmetry have been calculated. The relative stability of the isomeric Cu12Pt2 clusters follows the order: C3v > D2d > D4h. Each isomer has an active site for ethylene adsorption that consists of a single Pt atom surrounded by Cu atoms. The interaction of ethylene with the active site yields a π-C2H4 adsorption complex. The strongest π-C2H4 complex forms with the cluster of C3v symmetry; the bonding energy, ΔEπ(C2H4), is −15.6 kcal mol−1. The bonding energies for the π-C2H4 complex with Cu14 and Pt14 clusters are −6.5 and −18.8 kcal mol−1, respectively.The addition of Pt to Cu modifies the valence spd-band of the cluster as compared to a Cu14 cluster. The DOS near the Fermi level increases when C2H4 adsorbs on the Cu12Pt2 cluster. As well, the center of the d-band shifts toward lower binding energies. Ethylene adsorption also induces a number of states below the d-band. These states correspond to those of gas-phase C2H4.The vibrational frequencies of C2H4 adsorbed on the clusters of D4h and C3v symmetry have been calculated. The phonon vibrations occur below 250 cm−1. The intense bands around 200 cm−1 are attributed to stretching vibrations of the Pt-Cu bonds normal to the cluster surface. The stretching vibrations of the Pt-C bonds depend on the local structure of the active site: νs(Pt-C) = 268 cm−1 and νas(Pt-C) = 357 cm−1 for the cluster of the D4h symmetry; νs(Pt-C) = 335 cm−1 and νas(Pt-C) = 397 cm−1 for the cluster of the C3v symmetry. Bands in the range of 800-3100 cm−1 are attributed to vibrations of the adsorbed C2H4 molecule. The signature frequencies of the π-C2H4 adsorption complex are the δs(CH2) deformation vibration at ∼1200 cm−1 and the ν(C-C) stretching vibration at ∼1500 cm−1. These vibration are absent for di-σ-C2H4 adsorption complexes.  相似文献   

19.
The kinetics and mechanisms of the reactions of cyanomidyl radical (HNCN) with oxygen atoms and molecules have been investigated by ab initio calculations with rate constant prediction. The doublet and quartet state potential energy surfaces (PESs) of the two reactions have been calculated by single-point calculations at the CCSD(T)/6-311+G(3df, 2p) level based on geometries optimized at the CCSD/6-311++G(d, p) level. The rate constants for various product channels of the two reactions in the temperature range of 300-3000 K are predicted by variational transition state and RRKM theories. The predicted total rate constants of the O(3P) + HNCN reaction at 760 Torr Ar pressure can be represented by the expressions ktotal (O + HNCN) = 3.12 × 10−10 × T−0.05 exp (−37/T) cm3 molecule−1 s−1 at T = 300-3000 K. The branching ratios of primary channels of the O(3P) + HNCN are predicted: k1 for producing the NO + CNH accounts for 0.72-0.64, k2 + k9 for producing the 3NH + NCO accounts for 0.27-0.32, and k6 for producing the CN + HNO accounts for 0.01-0.07 in the temperature range studied. Meanwhile, the predicted total rate constants of the O2 + HNCN reaction at 760 Torr Ar pressure can be represented by the expression, ktotal(O2 + HNCN) = 2.10 × 10−16 × T1.28exp (−12200/T) cm3 molecule−1 s−1 at T = 300-3000 K. The predicted branching ratio for k11 + k13 producing HO2 + 3NCN as the primary products accounts for 0.98-1.00 in the temperature range studied.  相似文献   

20.
A series of continuous, crack-free, highly ordered amino-functionalized mesoporous silica thin films have been directly synthesized by co-condensation of tetraethoxysilane (TEOS) and 3-aminopropyltriethoxysilane (APTES) in the presence of cationic CH3(CH2)15N+(CH3)3Br (CTAB), nonionic C16H33(OCH2CH2)10OH (Brij-56) or triblock copolymer H(OCH2CH2)20(OCH(CH3)CH2)70(OCH2CH2)20)OH (P123) surfactant species under acidic conditions by sol-gel dip-coating. The molar ration of APTES/(TEOS + APTES) in the starting sol attains a value of 0.4. The effect of the sol aging on the mesostructure of thin films is systematically studied, and the optimal sol aging time is obtained for different surfactant systems. The amino-functionalized mesoporous silica thin films exhibit long-range ordering of 2D hexagonal (p6mm) and 3D cubic (Fm3m) pore arrays of size range from 2.2 to 8.3 nm following surfactants extraction as demonstrated by XRD, TEM and physical adsorption techniques. Based on BET surface area and weight loss, the surface coverage of amino-groups for thin films prepared using different surfactants is calculated to be 3.2 and above amino-groups per nm2, which is very useful and promising for incorporating inorganic ions and biomolecules into these mesoporous silica materials.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号